14 Comparison of DNA Replication in Cells from Prokarya and Eukarya

14 Comparison of DNA Replication in Cells from Prokarya and Eukarya Bruce Stillman Cold Spring Harbor Laboratory Cold Spring Harbor, New York 11724 I...
Author: Sydney Black
13 downloads 0 Views 2MB Size
14 Comparison of DNA Replication in Cells from Prokarya and Eukarya Bruce Stillman Cold Spring Harbor Laboratory Cold Spring Harbor, New York 11724

It is now supposed that utilization of DNA as genetic material emerged after an RNA world existed (Gesteland and Atkins 1993) and that all contemporary organisms evolved from a common ancestor related to primitive photosynthetic bacteria (Woese and Pace 1993). It is therefore not surprising that cells in the three major kingdoms, Bacteria, Archaea, and Eukarya, use roughly similar strategies and mechanisms for genome duplication. Nevertheless, the diversity of DNA replication is evident when the varied strategies used for replication of bacteriophage, plasmid, and virus genomes in both prokaryotes and eukaryotes are considered. This diversity becomes obvious when scanning the chapters that summarize the replication of virus genomes later in this book (Hassell and Brinton; Stenlund; Hay; Challberg; Yates; Traktman; Cotmore and Tattersall; Seeger and Mason; Bisaro; Ahrens et al.). I do not attempt to review these varied mechanisms here but rather focus on comparing the replication of cellular DNA in bacteria and eukaryotes. Unfortunately, except for some studies on DNA polymerases, little is known about the process of DNA replication in Archaea, and therefore inclusion of this kingdom adds little to the discussion. In organisms from bacteria to the multicellular eukaryotes, DNA replication is intimately coupled to the physiology of the individual cell and to global growth controls imposed on a population of cells, whether in a colony of bacteria or in a subset of cells within a developing organism. These regulatory pathways are different in bacteria and eukaryotes and also vary between different cell types within a single organism. Nevertheless, the actual process of DNA replication has striking parallels between the two groups of organisms. A universal challenge for the replication machinery is that the process must be precise and tightly controlled to ensure faithful transmission of the genetic heritage to the next generation of cells. If this goal is not achieved, disastrous consequences await the organism or its descendants. DNA Replication in Eukaryotic Cells 8 1996 Cold Spring Harbor Laboratory Press 0-87969-459-9/96 $5

+ .OO

435

436

6.Stillman

In general, the replication of chromosomes in both groups can be separated into multiple stages. First is the recognition of specific DNA sequences within the genome by initiator proteins. This interaction is the primary determinant of the location of the origin of DNA replication. At this location, a pre-replication complex of proteins is established that renders the genome competent for replication. The two important enzyme activities that must be attracted to the origin are a helicase that will eventually unwind the template DNA and a primase that will provide the necessary primer for initiation of DNA synthesis by DNA polymerases. Activation of the pre-replicative complex is a key step that eventually leads to the unwinding of the DNA and DNA synthesis. The establishment of a multiprotein complex called a replisome at the replication fork involves the assembly of the DNA polymerases and accessory proteins, followed by fork movement. Finally, termination of DNA replication occurs prior to separation of the daughter chromosomes. It is in each of these areas that the processes in bacteria and eukaryotes will be compared. This discussion relies heavily on the reviews presented in this volume on the replication mechanisms in eukaryotes for specific examples, and much of the primary literature is cited in these other chapters.

CHROMOSOME STRUCTURE AND REPLICATION

Perhaps the most obvious difference between the two classes of cells is how the genetic material is organized into chromosomes. Bacteria have a circular chromosome that contains one principal origin of DNA replication located at the OriC locus in Escherichia coli. The rate at which the DNA polymerases move along the E. coli chromosome is approximately 100 kb per minute, and thus the entire chromosome can be replicated from a single origin in about 40 minutes (Kornberg and Baker 1992). The genomes of eukaryotes are much larger than the bacterial chromosomes, and the rate of replication fork movement averages about 2 kb per minute, a rate that is prohibitive for even the smallest genome to replicate from a single origin (Fangman and Brewer 1992). Moreover, DNA in eukaryotes is invariably sequestered into multiple chromosomes, which demands more than one origin per genome. In certain artificial chromosomes in Saccharomyces cerevisiae, one origin is sufficient to replicate a chromosome of about 100 kb in size, but if the chromosome is larger than this limit, it becomes very unstable with only one origin (Wellinger and Zakian 1989; Dershowitz and Newlon 1993; Newlon, this volume). It has been known for more than three decades that different regions

Prokaryotic and Eukaryotic DNA Replication

437

of individual eukaryotic chromosomes replicate at distinct times during S phase, demonstrating that multiple origins of DNA replication must exist (Hand 1978; Fangman and Brewer 1992; DePamphilis, this volume). Unlike bacteria, eukaryotes initiate chromosome replication in a temporally regulated manner that can change during development (Simon and Cedar, this volume). The existence of multiple origins affords considerable flexibility in the timing of replication. In addition to the relative timing of replication of different regions of the genome, the total time taken to replicate the entire genome can vary considerably during development of an organism (Blumenthal et al. 1974; Callan 1974). This appears to be due to a change in the frequency of active origins of DNA replication distributed along the chromosome, rather than a change in the rate of replication fork movement. For example, in the rapid replication cycles of nuclei in the early Drosophila embryo, origins of DNA replication are located on average 3-7 kb apart, but later during embryogenesis, the frequency of replication origins changes to about 40-100 kb as cells begin to differentiate (Blumenthal et al. 1974). Interestingly, if mammalian cells are arrested for a long period of time at the G1 to S phase, the frequency along the genome of active origins of DNA replication increases following removal of the block, and the total time taken to pass through S phase is much shorter than the time taken in the absence of the inhibitor (Taylor 1977). This suggests that a factor may accumulate in the nucleus of cells that determines the frequency of active origins of DNA replication. A good candidate protein is the mammalian homolog of the protein from Schizosaccharomyces pombe and the related protein Cdc6p from S. cerevisiae (Kelly et al. 1993; Liang et al. 1995). A major difference between the two groups of cells emerges from these observations. The existence of a single origin in the bacterial chromosome fixes the relative order of gene duplication, except perhaps for occasional selective gene amplification. In contrast, eukaryotes have enormous flexibility in the relative time during S phase when specific genes replicate. This permits the replication of specific regions of the genome to be linked to gene expression, a topic discussed in an accompanying chapter (Simon and Cedar, this volume). This flexibility may be an important component of developmental patterning in cells of metazoan origin. Even though the number of active origins in a single bacterial chromosome does not vary, as described below in the section on control of DNA replication, the frequency of active origins within a cell can vary considerably in bacteria due to reinitiation in a single celldivision cycle.

438

B. Stillman

In bacteria that have distinct cell types, developmentally imposed control of the initiation of DNA replication can occur. In Caulobacter crescentus, cell division can produce two morphologically different cell types that have different replicative fates (Marczynski et al. 1995). One cell type, called the stalked cell, replicates the entire chromosome after cell division, whereas the other cell type, the swarmer cell, only replicates its DNA after a delay. In the stalked cell, a single chromosomal origin of DNA replication is utilized. The replicator element, Cori, has recently been characterized, and in addition to the initiator proteinbinding sites that are analogous to the DnaA initiator protein-binding sites, within the E. coli OriC replicator, Cori contains essential transcriptional promoter elements that are only active in the stalked cell (Marczynski et al. 1995). Thus, cell-type-specific transcription can influence replication. It has been known for some years that transcription can facilitate initiation of DNA replication from the bacteriophage h origin as well as the E. coli OriC (Kornberg and Baker 1992). The C. crescentus case is an example in bacteria where transcription may effect cell-type-specific initiation of DNA replication. Transcription factors in eukaryotic cells can also be intimately involved in the initiation of DNA replication and influence origin firing in a cell-type- and temporal-specific manner (van der Vliet, this volume; for review, see DePamphilis 1993). A recent example in mammalian cells is the locus controlling region (LCR) that regulates the developmentally controlled expression of the human pglobin gene. The LCR is required for the initiation of DNA replication from an origin located near the 0-globin gene (Aladjem et al. 1995). These selected examples demonstrate that despite the quite different organization of the chromosomes, and the number of replication origins, bacteria and eukaryotes can utilize similar strategies to control whether or not initiation of DNA replication will occur. Consideration of the influence of chromosome structure on DNA replication in bacteria and eukaryotes must also take into account the different organization of DNA in the cell. The bacterial chromosome is associated with the cell membrane but otherwise is exposed to the entire intracellular environment. In contrast, eukaryotic DNA replication occurs in a distinct compartment in the cell, affording the separation of proteins that may influence the initiation of DNA replication. Within the nucleus, initiation of eukaryotic DNA replication occurs at pre-replicative complexes that are established prior to the beginning of S phase, and these presumably contain the initiator and other replication proteins (Adachi and Laemmli 1994; Diffley et al. 1994). A similar initiation complex

Prokaryotic and Eukaryotic DNA Replication

439

may exist at the membrane-bound bacterial replicator, since DnaA protein from E. coli is a lipid-binding protein and is associated with the membrane (Sekimizu and Kornberg 1988; Sekimizu et al. 1988a,b). Thus, in both cell types, initiation may actually occur on a solid-state support, albeit that the supports and environment may be quite different.

REPLICATORS

The mechanism of initiation was first proposed for replication of the bacterial chromosome in the early 1960s (Jacob et al. 1964). In this classic paper, a cis-acting DNA element called the replicator was proposed to be a primary determinant of the location in the genome where initiation of DNA replication was to occur. Furthermore, it was posited that initiation of replication of the bacterial genome also required a trans-acting factor called the initiator that interacted with the replicator. The replicator can be defined by genetic means, and, indeed, the E. coli replicator OriC was characterized by the isolation of fragments from the E. coli genome that could support autonomous replication of recombinant plasmids (Hirota et al. 1979; Messer et al. 1979; von Meyenburg et al. 1979). These plasmids behaved in a manner similar to the complete chromosome. Fine-structure analysis by mutagenesis and evolutionary comparison of replicators from five enteric bacteria revealed that the bacterial replicator is highly conserved between species and is composed of multiple essential DNA sequence elements (Fig. 1A). The E. coli replicator has four recognition sites for the initiator protein DnaA that are flanked on one side by an A*T-rich sequence and on the other by a series of repeats that can be easily unwound in negatively supercoiled DNA (Kornberg and Baker 1992). These three repeats, or 13-mers, are locally unwound by DnaA once it is bound to OriC (Bramhill and Kornberg 1988). This general structure for bacterial replicators is also found in the origins of DNA replication from the lambdoid phages. Interspersed within the E. coli replicator are ten recognition sites for the DNA adenine methyltransferase (dam methylase), and methylation of these sites plays an important role in the control of initiation (see below). The replicators present in eukaryotic viruses have been characterized by mutational analysis. One of the best understood of these is the simian virus 40 (SV40) replicator called the SV40 ori (Stillman 1989; Hassell and Brinton, this volume). Like the bacterial replicators, the SV40 ori contains multiple binding sites for the SV40 initiator protein, T antigen, and these binding sites are flanked by an A*T-rich sequence on one side and a partial palindromic repeated sequence (Fig. 1B) (Parsons et al.

440

B. Stillman

A. OriC

DnaA

13-mer repeats

/ i\\ * Dam methylase

B. SV40 inverted repeat

I+

C. S. cerevisiae ORC

/I

ABF1

B3

82

B1

A

Figure 1 Comparison of the replicators from E. coli, SV40, and S. cerevisiae. (A) E. coli OriC. (B) The SV40 ori. (C) S. cerevisiae A R S l .

1990). The latter repeat also interacts with T antigen, resulting in the unwinding of 8 bp of DNA. Thus, the SV40 ori is remarkably similar in organization to the E. coli replicator. Understanding the nature of replicators in eukaryotic cell chromosomes has been more problematic, particularly in multicellular eukaryotes. Only in the yeasts have replicators been isolated by selection of fragments that permit autonomous replication of plasmids (Campbell and Newlon 1991; DePamphilis; Newlon; both this volume). Attempts to isolate such autonomous replicating sequences (ARSs) in other eukaryotes have not been successful, and thus the replicators from metazoan cells remain poorly defined. In S. cerevisiae, a number of replicators have been characterized extensively, and the general picture that has emerged is that the replicators have little similarity to the bacterial and viral replicators (Fig. 1C). The best-characterized replicator is ARSl from chromosome IV, and it consists of four DNA sequence elements, only one of which (A) is essential (Marahrens and Stillman 1992). The initiator protein for eukaryotes is the origin recognition complex (ORC), and this six-subunit protein binds to the A and B1 elements in an

Prokaryotic and Eukaryotic DNA Replication

441

ATP-dependent manner (Bell and Stillman 1992; Rao and Stillman 1995; Rowley et al. 1995). Another element, which is not essential and is only present in some replicators in the S. cerevisiae chromosomes, is a binding site for the transcription factor ABFl (Marahrens and Stillman 1992). Interestingly, other transcription factors can substitute for ABFl and activate initiation of DNA replication, similar to the stimulation of initiation of replication at virus origins by transcription factors (DePamphilis 1993; Hassell and Brinton; van der Vliet; both this volume). In general, the yeast replicators contain DNA sequences that are easily unwound when a plasmid in which they are cloned is in a highly negatively supercoiled state (Natale et al. 1993). A similar phenomenon has been observed with the E. coli OriC and the SV40 ori (Kowalski and Eddy 1989; Lin and Kowalski 1994), and in these cases, the region that unwinds first corresponds to the region that is unwound by the binding of DnaA or SV40 T antigen, respectively. This suggests that all origins of DNA replication have a tendency to unwind, and the energy for separation of the double helix either comes from torsional strain in the DNA, binding of the initiator protein, or both. For replication of plasmid DNAs containing the phage h or OriC replicators in vitro, torsional strain provided by a balance of gyrase, topoisomerase I, and the HU protein greatly facilitates origin unwinding and initiation of DNA replication (Sekimizu et al. 2988a,b; Mensa-Wilmot et al. 1989; Hwang and Kornberg 1992). Furthermore, transcription of the template DNA in the vicinity of the origin is necessary for the earliest stages of duplex strand unwinding during the initiation of DNA replication (Baker and Kornberg 1988; Learn et al. 1993). Whether transcription per se or the action of transcriptional activation domains present on sequence-specific DNA-binding proteins activates initiation at eukaryotic virus and cell origins remains to be determined.

INITIATORS

As indicated above, initiator proteins bind in a sequence-specific manner to the replicator, and together they function as the primary determinants of the location of the origin of DNA replication. The primary function of the initiator-replicator complex is as a landing pad for other replication proteins such as the helicase and the primase. Initiators can come in many different versions. All function as DNA-binding proteins, and all interact with other proteins that are required for the initiation of DNA replication. Some, such as the phage h 0 protein, function only in this capacity (Dodson et al. 1985), whereas others have additional activities.

442

B. Stillman

For example, the DnaA protein from E. cofi is an ATP/ADP-binding protein with the ability to slowly hydrolyze the ATP to ADP, whereas the SV40 T antigen, the papillomavirus E l protein, and the herpes simplex virus (HSV) UL9 initiators bind ATP, have a potent intrinsic ATPase activity, and in addition, contain DNA helicase function (Stahl et al. 1986; Sekimizu et al. 1988a,b; Bruckner et al. 1991; see Boroweic; Challberg; Hassell and Brinton; Stenlund; all this volume). One intriguing initiator from bacteriophage P4 is a sequence DNA-binding protein, an ATPase, a DNA helicase, and a DNA primase, and thus fulfills many of the functions required for initiation of DNA replication (Ziegelin et al. 1993). A common characteristic of initiator proteins is that they often oligomerize upon binding to the DNA. The DnaA and h 0 initiator proteins form complexes of 8 or about 20 proteins bound to OriC or orih, respectively, even though there are only four DNA recognition sequences in the replicator (Echols 1986; Kelman and O'Donnell 1994). Similarly, the SV40 T antigen has four recognition sites in the replicator, yet two hexamers of T antigen bind to the SV40 ori in an ATP-dependent manner (Dean et al. 1987). The oligomerization probably acts to stabilize the initiator-replicator interactions and to allow efficient interaction with other proteins. In the case of SV40 T antigen, the functional helicase unit is a hexamer. In many respects, the T-antigen hexamer resembles the RuvB ATPase that is involved in branch migration of DNA at a recombination Holliday junction (Parsons et al. 1995). Oligomerization of an initiator protein can prevent DNA replication. The bacteriophage P1 genome can exist in a lysogenic state in its host bacteria as a plasmid, and initiation of DNA replication requires the OriPZ replicator and the plasmid-encoded RepA initiator (Baker and Wickner 1992). The RepA protein exists as a stable dimer when not bound to the DNA. The E. coli chaperone proteins DnaJ, DnaK, and GrpE combine to break apart the RepA dimer into RepA monomers, thereby allowing RepA to bind to OriPZ in cooperation with the DnaA protein to form a productive replicator-initiator complex (Baker and Wickner 1992). In many cases, but not all, the binding of the initiator protein to the replicator causes local unwinding of a region of DNA at the origin of DNA replication. The DnaA protein induces unwinding of the 13-mer elements at OriC, and SV40 T antigen causes unwinding of 8 bp in the partial "early" palindrome within the SV40 ori (Bramhill and Kornberg 1988; Borowiec et al. 1990; Hassell and Brinton; Brush and Kelly; both this volume). Unwinding of the DNA at the origin of DNA replication is a key step in the pathway leading to initiation of DNA synthesis at the

Prokaryotic and Eukaryotic DNA Replication

443

origin. This is facilitated by regions within the origin of DNA replication that have a low helical stability and thus have a tendency to unwind when placed under torsional strain (Kowalski and Eddy 1989; Natale et al. 1993; Lin and Kowalski 1994). The assembly and disassembly of the multiprotein complex on the replicator and the ability of the complex to function as a landing pad are highly regulated processes. The function of the E. coli DnaA complex is regulated by membrane-bound acidic phospholipids (Crooke et al. 1992). The ATP form, but not the ADP form, of DnaA is able to unwind the 13mers in OriC, and membrane-bound acidic phospholipids greatly stimulate the exchange of ADP to ATP bound to DnaA. Exchange to the active ATP-DnaA complex requires the DnaA to be bound to OriC because the acidic phospholipids inactivate DNA-free DnaA protein so that it cannot bind OriC.These results suggest a link between DnaA activity and the membrane localization of OriC. Phosphorylation of proteins is not known to be a direct mechanism for control of DNA replication in bacteria, but it is a major mechanism for regulation of some eukaryotic virus initiator proteins. For example, the ability of SV40 T antigen to unwind the origin of DNA replication is positively regulated by phosphorylation of a single threonine by a cyclindependent protein kinase (Ismail et al. 1993; McVey et al. 1993). On the other hand, the DNA-binding activity is negatively regulated by phosphorylation of two serine residues by casein kinase I, which can be reversed by the action of phosphatase 2A (Cegielska et al. 1994a,b). In eukaryotes, phosphorylation can also control the cellular localization of proteins. For example, phosphorylation by casein kinase I1 of serine residues that lie adjacent to the nuclear localization signal in SV40 T antigen greatly facilitates the transport of the protein from the cytoplasm to the nucleus. It is easy to imagine that this may play a regulatory role for the initiation of DNA replication. Thus, in eukaryotes, unlike bacteria, control of the intracellular locale of proteins may affect their function. A discussion of the role of the nuclear envelope in the initiation of cellular DNA replication can be found in Laskey and Madine (this volume).

UNWINDING AT THE ORIGIN

Following DNA sequence recognition and establishment of the initiatorprotein complex at the origin of DNA replication, the next important events are unwinding of the template DNA and priming DNA replica-

444 B. Stillman

tion. Initiator proteins such as DnaA protein (Bramhill and Kornberg 1988) and SV40 T antigen (Borowiec et al. 1990) can locally unwind a region of the origin of DNA replication, and this region of singlestranded DNA is presumed to be the entry site for a DNA helicase and single-stranded DNA-binding proteins such as the E. coli DnaB helicase and SSB, respectively. A DNA helicase has to be loaded onto the initiator-DNA complex for more extensive unwinding to occur. SV40 T antigen is an unusual case because it is a DNA helicase as well as an initiator protein; thus, the helicase is automatically located in a prime site for extensive unwinding. Indeed, T antigen unwinds almost the entire SV40 genome in cooperation with the eukaryotic cell single-stranded DNA-binding protein RP-A and a topoisomerase (Borowiec et al. 1990; see Borowiec; Hassell and Brinton; both this volume). In most cases, however, the helicase is a separate polypeptide from the initiator protein and needs to be loaded onto the DNA. Loading of the DNA helicase is a complicated affair. The DnaB helicase from E. coli is used to initiate DNA replication from OriC and is also required for replication of the bacteriophage h genome (Baker and Wickner 1992; Stillman 1994a). In neither case can the DnaB helicase recognize the initiator protein (DnaA or h 0 protein) by itself, but needs to be chaperoned by a specialized protein. The E. coli DnaC and the phage h P proteins function as helicase-loading proteins because they both bind to the DnaB helicase and also bind either to the DnaA or h 0 proteins, respectively (Fig. 2). These helicase-loading proteins can inhibit the DNA helicase activity, and it is necessary to remove them before the helicase can be activated to unwind the template DNA (Stillman 1994a). For initiation at the phage h origin (orih), the E. coli heat shock chaperone proteins DnaJ, DnaK, and GrpE cooperate to remove h P protein and thus activate the DnaB helicase (Baker and Wickner 1992; Stillman 1994a). Here is another example from the prokaryotic world where chaperone proteins pull apart protein complexes that are involved in the initiation of DNA replication. To date, chaperones have not been shown to play a role in disassembling protein complexes in eukaryotes, but little is known about the eukaryotic pre-replicative complex. It is likely that similar assembly and disassembly of multiprotein complexes occur at eukaryotic replicators. The helicase-loading proteins can also facilitate the assembly of a DNA helicase onto the template DNA in the absence of an origin of DNA replication. For example, the bacteriophage T4 gene 59 protein (gp59) binds to both single-stranded DNA and the phage single-strandbinding protein, gp32 (Barry and Alberts 1994; Morrical et al. 1994).

Prokaryotic and Eukaryotic DNA Replication

4

445

0 protein

& &B

4 4

+

protein helicase

Removal of P protein by DnaJ, DnaK, GrpE

c PRIMING AND INITIATION OF DNA REPLICATION

Figure 2 Loading of a DNA helicase. The bacteriophage h P protein loads the DNA helicase onto the DNA. The h P protein must be removed before the helicase (DnaB) can function to unwind the DNA prior to DNA replication. This is facilitated by the chaperone proteins DnaJ, DnaK, and GrpE. (Reprinted, with permission, from Stillman 1994a.)

When gp32 coats a single-stranded DNA template, the phage helicase gp41 cannot bind to the DNA. The gp59 protein loads the gp41 helicase onto the gp32-coated DNA,and the helicase can now move rapidly along the template in search of duplex DNA to unwind. This mechanism of helicase loading is almost certainly used during the elongation stages of DNA replication. In a similar manner, SV40 T antigen can bind the RPA protein, the eukaryotic SSB, facilitating loading of a helicase onto the RP-A-coated single-stranded DNA (Collins and Kelly 1991; Melendy and Stillman 1993). Although many DNA helicases have been identified in eukaryotes, it is not clear which are required for the initiation of DNA replication (Borowiec, this volume). A good candidate is the helicase B from a murine cell line that harbors a temperature-sensitive helicase which causes a defect in DNA replication at the nonpermissive temperature (Seki et al. 1995). Moreover, although proteins such as Cdc6p that interact with the eukaryotic initiator protein ORC are beginning to be identified (Liang et al. 1995), it is not yet clear whether the same mechanism of helicase loading occurs in eukaryotic cells as occurs in bacteria.

446

B. Stillman

PRIMING DNA SYNTHESIS

Following unwinding of the template DNA surrounding the replicator, priming of DNA synthesis occurs, usually by a DNA primase. As discussed by others in this volume (Brush and Kelly; Salas et al.), priming can occur by a number of different mechanisms, including priming by a nucleotide linked to a protein such as occurs during @29 phage and adenovirus DNA replication. For initiation of chromosomal DNA replication in bacteria and almost certainly in eukaryotes, priming involves the assembly of a primase on the DNA. In E. coli,the primase exists as a free enzyme that is not always bound to the DNA polymerase (Baker and Wickner 1992; Kornberg and Baker 1992), whereas in eukaryotes, the primase enzyme is always in a complex with DNA polymerase-a (Wang, this volume). Herein lies a significant difference between prokaryotes and eukaryotes: Bacteria contain only one principal replicative DNA polymerase (DNA polymerase 111 in E. coli) that must function on both the leading and lagging strands at a replication fork, but in eukaryotes, the DNA polymerizing activity is shared by three DNA polymerases, pol-a, pol-6, and pol-&(Kornberg and Baker 1992; Wang, this volume). Thus, the pol 111 enzyme in E. coli needs to be free of the primase when replicating the continuously synthesized leading strand, but must cyclically interact with the primase during lagging-strand replication (Hacker and Alberts 1994a,b; Stillman 1994b; Stukenberg et al. 1994). Eukaryotes have devised a DNA polymerase switching mechanism that involves the pol-a:primase complex in priming DNA replication at the origin of DNA replication and for every Okazaki fragment, and then a second DNA polymerase that is not associated with the primase (pol-6 or E ) completing the synthesis (Figs. 3 and 4) (Nethanel and Kaufmann 1990; Tsurimoto et al. 1990; Waga and Stillman 1994; Stillman 1994b). A key question is how the primase finds either the origin of DNA replication or the site at which Okazaki fragments must begin. Bacterial and phage primases can exist in two different classes: those that are part of the helicase polypeptide and others that exist as proteins separate from the helicase polypeptide (Ilyina et al. 1992). If the primase activity coexists in the same polypeptide as the helicase, then the primase is associated with the replication fork by default. For bacterial chromosome replication and for the replication of some eukaryote virus genomes such as SV40 and herpesvirus, the primase-loading problem has been solved by the primase having an affinity either for the helicase that has already been loaded onto the template DNA or, alternatively, for a structure in the DNA created by the helicase (Baker and Wickner 1992; Marians

Prokaryotic and Eukaryotic DNA Replication

447

3‘ 5’

- p I PCNA 5‘

3’

Figure 3 Replication fork proteins. Model for a dimeric DNA polymerase at the replication fork. (Reprinted, with permission, from Stillman 1994b [copyright Cell Press].)

1992; Kelman and O’Donnell 1994; Stillman 1994a; Hassell and Brinton; Challberg; both this volume). For example, the SV40 T antigen binds directly to the pol-a:primase complex and can load the primase onto an RP-A-coated template DNA by direct interaction with RP-A (Collins and Kelly 1991; Melendy and Stillman 1993; Dornreiter et al. 1993). In E. coli, the DnaB helicase activates the DnaG primase (Baker and Wickner 1992; Marians 1992). Once located at the priming site, the primase in both prokaryotes and eukaryotes synthesizes a short (8-12 nucleosides) RNA primer that is utilized by the DNA polymerase for synthesis of DNA.

ASSEMBLY OF THE DNA POLYMERASE HOLOENZYME

A common feature of the DNA replication fork in bacteria and eukaryotes is that DNA is synthesized in a semidiscontinuous manner, with one strand, the leading strand, polymerized in a continuous manner and the other lagging strand formed in a discontinuous manner via the synthesis of Okazaki fragments (Brush and Kelly, this volume). The DNA polymerase is loaded onto the template DNA at the site of the primer by a mechanism that is remarkably conserved between prokaryotes and eukaryotes (Table 1) (Kelman and O’Donnell 1994; Stillman 1994b). The process is accomplished by polymerase accessory

-

.d

0 0

cri

Prokaryotic an d Eukaryotic DNA Replication

449

Table 1 Functions of DNA replication fork proteins E. coli

Phage T4

SV40Ihuman

Functions

DnaB

41

T antigen

DnaC

59

T antigen

SSB

32

RP-A

y-Complex (Y 6.6 '

44/62

RF-C

z

43 45

? PCNA

pol 111 core (see) DnaG

43

pols

61

DNA helicase; stimulates priming on ssDNA allows loading of helicase and primase onto SSB-coated DNA (primosome assembly) single-stranded DNA-binding; stimulates DNA polymerase; facilitates helicase loading DNA-dependent ATPase; primertemplate binding; stimulates DNA polymerase dimerization of holoenzyme stimulates DNA polymerase; stimulates DNA-dependent ATPase DNA polymerase 3 ' +5 ' exonuclease primase

Ligase DNA pol I

T4 ligase 43

RNase H

RNase H

primase pol-ab ligase I ligation of DNA FEN-l/MFl nuclease for removal of RNA primers nuclease for removal of RNA RNase H1 primers

xw)

P

(Reprinted, with permission, from Stillman 1994b [copyright Cell Press].) aDNA polymerase-6 has been shown to function in SV40 DNA replication; an essential DNA polymerase, DNA polymerase+, has not yet been assigned a specific function in DNA replication. *he human DNA polymerase-a and primase activities function as a multiprotein complex to synthesize RNADNA primers. (RP-A) Replication protein A; (RF-C) replication factor C; (PCNA) proliferating cell nuclear antigen; (SSB) single-strand DNA-binding protein.

proteins. The eukaryotic proteins are discussed by Hiibscher et al. (this volume). Similar mechanisms and enzymatic activities in these eukaryote proteins and the bacteriophage T4 counterparts were recognized first (Tsurimoto and Stillman 1990; Tsurimoto et al. 1990), and later, some sequence similarities were observed (O'Donnell et al. 1993). The polymerase accessory proteins from E. cofi, bacteriophage T4, and eukaryotes include the single-strand-binding protein (SSB, gp32, or RP-A), a polymerase clamp protein (p subunit of pol 111, gp45, or PCNA), and a DNA-dependent ATPase that loads the polymerase clamp onto the DNA (y66'zxw, gp44/62, and RF-C). Only in the case of the

450

B. Stillman

clamp-loading proteins have significant DNA sequence similarities been observed (O'Donnell et al. 1993), but there is remarkable structural similarity between the polymerase clamp proteins even though there is no primary amino acid sequence similarity (Fig. 5) (Krishna et al. 1994). The p subunit of the pol I11 enzyme from E. coli and the S. cerevisiae PCNA have three-dimensional structures that can be almost superimposed one on top of the other. Both form a donut or torus-like shape, and the DNA passes through the hole in the middle of the structure. The fJ subunit contains 366 amino acids, whereas the PCNA contains only 258 amino acids, approximately two-thirds the size of the bacterial counterpart. Two monomers of the fi subunit form the torus structure, whereas three monomers of the smaller PCNA combine to form an identical structure. Each torus ring has an internal sixfold symmetry. The 228-amino acid phage T4 gp45 protein is presumed to also form a trimer. Interestingly, the carrot plant has two highly related PCNA proteins that are similar in sequence to PCNA from human and yeast; one is 264 amino acids, similar to the "eukaryotic short" form and the other is 365 amino acids, similar to the "bacterial long" form (Kelman and O'Donnell 1995). It has been suggested that these plants have one dimer PCNA and one trimer PCNA that may function at different times during development when DNA replication rates vary. Alternatively, since PCNA also functions in DNA repair (Nichols and Sancar 1992; Shivji et al. 1992), it is possible that one PCNA is involved in DNA repair and the other in DNA replication.

Polymerase 111 p subunit

Proliferatina Cell Nuclear Antiaen

Figure 5 Comparison of the structures of the clamp proteins from bacteria and yeast. Both proteins form a torus or donut-shaped molecule with sixfold symmetry. The E. coli pol 111 p subunit forms a dimer and PCNA forms a trimer. The arrows represent a repeated structural motif in the proteins. The DNA passes through the center of the structure.

Prokaryotic and Eukaryotic DNA Replication

451

The mechanism of polymerase loading is very similar in bacteria and eukaryotes and is discussed in detail elsewhere in this volume (Hiibscher et al.) and in other reviews (Kelman and O’Donnell 1994; Stillman 1994b). In both systems, an ATP-dependent clamp-loading protein recognizes a primer-template junction and assembles the polymerase clamp onto the DNA. The single-stranded DNA-binding protein that coats the single-stranded template DNA facilitates the recognition of the primer-template junction by preventing the clamp-loading protein from binding to single-stranded DNA. Once the clamp complex is assembled, a DNA polymerase is assembled onto the primer ready for DNA synthesis.

TOPOLOGY AND DNA REPLICATION

Topoisomerases in both prokaryotes and eukaryotes deal with the topological problems associated with unwinding the double helix during DNA replication and with separating the daughter chromosomes after DNA replication is completed. The different types of topoisomerase have been reviewed recently (Roca 1995), and a discussion of the eukaryotic topoisomerases appears in Hangaard Andersen et al. (this volume). The role of topoisomerases from bacteria and eukaryotes in DNA replication has been well discussed in a recent review (Ullsperger et al. 1995), and only some general points are noted here. Bacterial chromosome topology is controlled by a balance between topoisomerases that relieve torsional strain on the one hand and gyrases that induce negative supercoiling on the other. There is net negative supercoiling in the bacterial chromosome, and these negative supercoils are available to facilitate unwinding of the double helix for processes such as transcription and the initiation of DNA replication. In contrast, eukaryotic chromosomes have more complex topology. Eukaryotic chromosomal DNA has the same supercoil density as found in bacteria, but all the negative supercoils are believed to be constrained in nucleosomes (Ullsperger et al. 1995). Furthermore, the eukaryotic chromosome is folded into complex higher-order structures that the DNA replication apparatus must deal with. Thus, the DNA replication machineries in the two cell types must deal with the very different templates, although as indicated above, the basic replication proteins are quite similar. Unwinding of the DNA helix during replication fork progression generates positive supercoils ahead of the replication fork in the unreplicated DNA. There are different mechanisms to remove these positive supercoils (Ullsperger et al. 1995). The first is to actively generate negative

B. Stillman

452

Table 2 Topoisomerase activities topo I

Name

Type Genes (+) Supercoil removal in

vivo Decatenation Relaxesonly

Prokaryotes topo TI topo 111 topo TV

Eukaryotes topo I topo I1 topo TI1

-

-

-

-

-

1 topA

2 gyrA, gYrB

1 topB

1 top1

2 top2

1 top3

no

yes

no

2 parC, parE yes?

yes

yes

no

no yes

yes no

no

yes no

no no

yes no

no Yes

omega gyrase

yes

(-) supercoils (Adapted from Ullsperger et al. 199.5 [copyright Springer-Verlag GmbH].)

supercoils in the unreplicated DNA before the positive supercoils accumulate, a role for the bacterial type I1 topoisomerase commonly referred to as DNA gyrase (Table 2). The bacterial gyrase can keep up with the generation of positive supercoils ahead of the replication fork by introducing negative supercoils. A second mechanism is to remove directly the positive supercoils in the unreplicated DNA; both the eukaryotic type I and type I1 topoisomerases can perform this function (Table 2). This mechanism is thought to occur in the early stages of replication from an origin when there is a relatively large amount of unreplicated DNA. In the later stages of DNA replication, it is more difficult to introduce positive supercoils into the unreplicated DNA, and the replicated DNA tends to become interwound or catenated (Ullsperger et al. 1995). Thus, a third mechanism for relieving the positive supercoils generated by replication fork progression is to remove catenated DNA, a job for different types of topoisomerases in bacteria and eukaryotes. The bacterial type I topoisomerase, top0 111, binds to single-stranded DNA and thus can operate behind the replication fork to decatenate the replicated DNA. In contrast, the type I1 topoisomerases, top0 I1 in eukaryotes and type IV in bacteria, can decatenate the replicated duplex DNA and thus segregate completely replicated daughter chromosomes (Ullsperger et al. 1995). Finally, as pointed out by Ullsperger et al. (1995), if the replication fork contains a dimeric DNA polymerase with physically linked polymerase molecules of the type shown in Figure 3, this creates even more problems with DNA topology because the DNA polymerases are

Prokaryotic and Eukaryotic DNA Replication

453

not free to rotate about the template DNA. If the DNA polymerases are connected to each other, as seems likely in bacteria and eukaryotes, then enzymatic mechanisms must exist that deal with the passage of the double helix through the polymerase complex.

FIDELITY AND DNA REPAIR

Both bacteria and eukaryotic cells have mechanisms for repairing DNA (Friedberg et al. 1995), and it is beyond the scope of this discussion to review these varied mechanisms here (for eukaryotes, see Roberts and Kunkel; Friedberg and Wood; both this volume). As discussed by Nasmyth (this volume) and others elsewhere (Li and Deshaies 1993; Murray 1994; Nurse 1994), eukaryotic cells have a monitoring system that links DNA replication with cell-cycle progression, cell growth, and DNA repair. The existence of DNA damage-checkpoint mechanisms in bacteria that temporally delay DNA replication and cell-cycle progression in response to DNA damage is not as clear as has become apparent in eukaryotes, but inhibition of DNA synthesis and recovery are known to occur (Bridges 1995). Although the mechanisms of DNA repair are similar between bacteria and eukaryotes, the mechanisms that control the cell division cycle are very different. Thus, simple parallels between prokaryotes and eukaryotes have not emerged.

CONTROL OF INITIATION

Because the pathways that control the cell division cycle are very different between bacteria and eukaryotes, the mechanisms for controlling the frequency of initiation are also different. Methylation of the bacterial origin of DNA replication by the dam methylase causes the bacterial DNA to become hemi-methylated immediately after it has replicated (Baker 1994). In the E. coli OriC replicator, the Dam methylation sequences (5 ' -GATC-3 ' ) are clustered in the 13-mer repeats that unwind in the presence of DnaA and ATP. OriC is negatively regulated by a process called sequestration. For some time after the initiation of DNA replication, the hemi-methylated DNA is unable to reinitiate, thereby limiting multiple rounds of DNA replication in a cell. The product of the SeqA gene is required for sequestration, and SeqA interacts genetically with DnaA (Campbell and Kleckner 1990; Baker 1994). Recently, it has been shown that the SeqA protein binds to hemi-methylated DNA from either the E. coli OriC 13-mer region or from the bacteriophage P1 origin, PloriR (Brendler et al. 1995). Binding to these replication origins

454

B. Stillman

is preferred because these sequences are somehow slow to be fully methylated and therefore remain in the hemi-methylated state for some time. These data suggest that the SeqA protein might prevent initiation of DNA replication by interacting with the newly hemi-methylated replicator, thereby preventing productive binding of the replicator by the initiator DnaA. Methylation probably does not play a role in the regulation of initiation of DNA replication in eukaryotes. More likely, a multiprotein complex is formed on the DNA at the site of the replicator to produce a "competent" pre-replicative complex. This complex can then be acted on by other regulators of DNA replication that coordinate the initiation of DNA replication with progression through the cell division cycle (see Nasmyth, this volume). In many ways, this type of regulation may be similar to the assembly and disassembly of proteins at the phage h replicator (Stillman 1994a). This type of regulation may also facilitate the coordination of replication from multiple origins of DNA replication that does not occur in bacteria. TERMINATION OF DNA REPLICATION

The termination of DNA replication has been well studied in bacteria, including E. coli and Bacillus subtilis (Baker 1995). Replication of the bacterial chromosome terminates when the two replication forks that diverged from OriC converge in a region of genome that is 180° from the origin. There are multiple cis-acting sites in the DNA called Ter sites in E. coli and imperfect repeats (IR-I and IR-11) in B. subtilis. These DNA elements have directionality with respect to the advancing replication fork and interact with a protein called Tus in E. coli and RTP in B. subtilis. The Ter-Tus protein complex functions as an anti-helicase complex because it blocks the progression of DNA unwinding by a DNA helicase in the absence of DNA replication. The Tus-Ter complex has been shown to block the DnaB helicase (Khatri et al. 1989; Lee et al. 1989) and, interestingly, it can also inhibit the SV40 T-antigen helicase (Bedrosian and Bastia 1991). A more thorough description of termination is given by Bastia and Mohanty (this volume). Replication termination in eukaryotes is not well studied, but it is clear that specific DNA sequences are not generally required. For example, the SV40 origin of DNA replication has been cloned into many plasmids, and these plasmids replicate very well when they are introduced into primate cells, yet the DNA sequences located in the region opposite the origin are very different (Lai and Nathans 1975). Similarly, replica-

Prokaryotic and Eukaryotic DNA Replication

455

tion forks that meet each other following initiation at adjacent origins in eukaryotic chromosomes are believed to terminate when the two forks meet each other, but in a sequence-independent manner. There are, however, DNA sequences in eukaryotic genomes that cause DNA replication to arrest and occasionally terminate. For example, repeated DNA recognition sites for the Epstein-Barr virus (EBV) nuclear antigen 1 (EBNA1) cause replication forks to pause or arrest completely (Little and Schildkraut 1995). Furthermore, DNA replication fork barriers have been found in the ribosomal DNA repeats in s. cerevisiae (Brewer et al. 1992). Thus, it is possible that sequence-specific DNAbinding proteins may affect the passage of the eukaryotic DNA replication fork in much the same way as the Tus-Ter system in E. coli and that these sites may play some regulatory role that has yet to be appreciated. REFERENCES

Adachi, Y.and U.K. Laemmli. 1994. Study of the cell cycle-dependent assembly of the DNA prereplication centres in Xenopus egg extracts. EMBO 13: 4153-4164. Baker, T.A. 1994. A new controller in Escherichia coli. Curr. Opin. Genet. Dev. 4: 945-946. Aladjem, M.I., M. Groudine, L.L. Brody, E.S. Dieken, R.E.K., Fournier, G.M. Wahl, and E.M. Epner. 1995. Participation of the human P-globin locus control region in initiation of DNA replication. Science 270: 815-819. -. 1995. Replication arrest. Cell 80: 521-524. Baker, T.A. and A. Kornberg. 1988. Transcriptional activation of initiation of replication from the E. coli chromosomal origin: An RNA-DNA hybrid near oriC. Cell 55: 113-123. Baker, T.A. and S.H. Wickner. 1992. Genetics and enzymology of DNA replication in Escherichia coli. Annu. Rev. Genet. 26: 447-477. Barry, J. and B. Alberts. 1994. Purification and characterization of bacteriophage T4 gene 59 protein. J. Biol. Chem. 269: 33049-33062. Bedrosian, C.L. and D. Bastia. 1991. Escherichia coli replication terminator protein impedes simian virus 40 (SV40) DNA replication fork movement and SV40 large tumor antigen helicase activity in vitro at a prokaryotic terminus sequence. Proc. Natl. Acad. Sci. 88: 2618-2622. Bell, S.P. and B. Stillman. 1992. ATP-dependent recognition of eukaryotic origins of DNA replication by a multiprotein complex. Nature 357: 128-134. Blumenthal, A.B., H.J. Kriegstein, and D.S. Hogness. 1974. The units of DNA replication in Drosophila melanogaster chromosomes. Cold Spring Harbor Syrnp. Quant. Biol. 38: 205-223. Borowiec, J.A., F.B. Dean, P.A. Bullock, and J. Hurwitz. 1990. Binding and unwinding-How T-antigen engages the SV40 origin of DNA replication. Cell 60: 181-184. Bramhill, D. and A. Kornberg. 1988. A model for initation at origins of DNA replication. Cell 54: 915-918. Brendler, T., A. Abeles, and S. Austin. 1995. A protein that binds to the P1 origin core and the oriC 13-mer region in a methylation-specific fashion is the product of the host seqA gene. EMBO J. 14: 4083-4089.

456

B. Stillman

Brewer, B., D. Lockshon, and W. Fangman. 1992. The arrest of replication forks in the rDNA of yeast occurs independently of transcription. Cell 71: 267-276. Bridges, B.A. 1995. Are there DNA damage checkpoints in E. coli? BioEssays 17: 63-70. Bruckner, R.C., J.J. Crute, M.S. Dodson, and I.R. Lehman, 1991. The herpes simplex virus origin binding protein: A DNA helicase. J. Biol. Chem. 266: 2669-2674. Callan, H.G. 1974. DNA replication in the chromosomes of eukaryotes. Cold Spring Harbor Symp. Quant. Biol. 38: 195-203. Campbell, J.L. and N. Kleckner. 1990. E coli oriC and the dnaA gene promoter are sequestered from dam methyltransferase following the passage of the chromosomal replication fork. Cell 62: 967-979. Campbell, J.L. and C.S. Newlon. 1991. Chromosomal DNA replication. In The molecular and cellular biology of the yeast Saccharomyces: Genome dynamics, protein synthesis, and energetics (ed. J.R. Broach), pp. 41-146. Cold Spring Harbor Laboratory Press. Cold Spring Harbor, New York. Cegielska, A., I. Moarefi, E. Fanning, and D.M. Virshup. 1994a. T-antigen kinase inhibits simian virus 40 DNA replication by phosphorylation of intact T antigen on serines 120 and 123. J. Virol. 68: 269-275. Cegielska, A., S. Shaffer, R. Derua, J. Goris, and D.M. Virshup. 1994b. Different oligomeric forms of protein phosphatase 2A activate and inhibit simian virus 40 DNA replication. Mol. Cell. Biol. 14: 4616-4623. Collins, K.L. and T.J. Kelly. 1991. Effects of T antigen and replication protein A on the initiation of DNA synthesis by DNA polymerase a-primase. Mol. Cell. Biol. 11: 21 08-21 15. Crooke, E., C. Castuma, and A. Kornberg. 1992. The chromosome origin of Escherichia coli stabilizes DnaA protein during rejuvenation by phospholipids. J. Biol. Chem. 267: 16779-16782. Dean, F.B., M. Dodson, H. Echols, and J. Hurwitz. 1987. ATP-dependent formation of a specialized nucleoprotein structure by simian virus 40 (SV40) large tumor antigen at the SV40 replication origin. Proc. Natl. Acad. Sci. 84: 8981-8985. DePamphilis, M.L. 1993. How transcription factors regulate origins of DNA replication in eukaryotic cells. Trends Cell Biol. 3: 1161-163. Dershowitz, A. and C.S. Newlon. 1993. The effect on chromosome stability of deleting replication origins. Mol. Cell. Biol. 13: 391-398. Diffley, J.F.X., J.H. Cocker, S.J. Dowell, and A. Rowley. 1994. Two steps in the assembly of complexes at yeast replication origins in vivo. Cell 78: 303-316. Dodson, M., J. Roberts, R. McMacken, and H. Echols. 1985. Specialized nucleoprotein structures at the origin of replication of bacteriophage h: Complexes with hO protein and with LO, LP, and Escherichia coli DnaB proteins. Proc. Natl. Acad. Sci. 82: 4678-4682. Dornreiter, I., W.C. Copeland, and T.S.-F. Wang. 1993. Initiation of simian virus 40 DNA replication requires the interaction of a specific domain of human DNA polymerase a with large T antigen. Mol. Cell. Biol. 13: 809-820. Echols, H. 1986. Bacteriophage W development: Temporal switches and the choice of lysis or lysogeny. Trends Genet. 2: 26-30. Fangman, W. and B. Brewer. 1992. A question of time: Replication origins of eukaryotic chromosomes. Cell 71: 363-366. Friedberg, E.C., G.C. Walker, and W. Siede. 1995. DNA repair and mutagenesis. ASM

Prokaryotic and Eukaryotic DNA Replication

457

Press, Washington, D.C. Gesteland, R.F. and J.F. Atkins, eds. 1993. The RNA world. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York. Hacker, K.J. and B.M. Alberts. 1994a. The rapid dissociation of the T4 DNA polymerase holoenzyme when stopped by a DNA hairpin helix. J. Biol. Chem. 269: 24221-24228. 1994b. The slow dissociation of the T4 DNA polymerase holoenzyme when stalled by nucleotide omission. J. Biol. Chem. 269: 24209-24220. Hand, R. 1978. Eucaryotic DNA: Organization of the genome for replication. Cell 15: 317-325. Hirota, Y., S. Yasuda, M. Yamada, A. Nishimura, K. Sugimoto, H.Sugisaki, A. Oka, and M. Takanami. 1979. Structural and functional properties of the Escherichia coli origin of DNA replication. Cold Spring Harbor Symp. Quant. Biol. 43: 129-138. Hwang, D. and A. Kornberg. 1992. Opening of the replication origin of Escherichiu coli by DnaA protein with protein HU or IHF. J. Biol. Chem. 267: 23083-23086. Ilyina, T.V., A.E. Gorbalenya, and E.V. Koonin 1992. Organization and evolution of eubacterial and bacteriophage primase-helicase systems. J. Mol. Evol. 34: 351-357. Ismail, F., D. Small, 1. Gilbert, M. Hopfner, S. Randall, C. Schneider, A. Russo, U. Ramsperger, A. Arthur, H. Stahl, T. Kelly, and E. Fanning. 1993. Mutation of the cyclindependent kinase phosphorylation site in simian virus 40 (SV40) large T antigen specifically blocks SV40 origin DNA unwinding. J. Virol. 67: 4992-5002. Jacob, F., S. Brenner, and F. Cuzin. 1964. On the regulation of DNA replication in bacteria. Cold Spring Harbor Symp. Quant. Biol. 28: 329-348. Kelly, T.J., G.S. Martin, S.L. Forsburg, R.J. Stephen, A. Russo, and P. Nurse. 1993. The fission yeast cdcl8+ gene product couples S. phase to START and mitosis. Cell 74: 371 -382. Kelman, Z. and M. O’Donnell. 1994. DNA replication: Enzymology and mechanisms. Curr. Opin. Genet. Dev. 4: 185-195. -. 1995. Embryonic PCNA: A missing link? Curr. Biol. 5: 814. Khatri, G.S., T. MacAllister, P.R. Sista, and D. Bastia. 1989. The replication terminator protein of E. coli is a DNA sequence-specific contra-helicase. Cell 59: 667-674. Kornberg, A. and T.A. Baker. 1992. DNA replication, 2nd edition. W.H. Freeman, New York. Kowalski, D. and M.J. Eddy. 1989. The DNA unwinding element: A novel, cis-acting component that facilitates opening of the Escherichia coli replication origin. EMBO J. 8: 4335-4344. Krishna, T.S.R., X.-P. Kong, S. Gary, P.M. Burgers, and J. Kuriyan. 1994. Crystal structure of the eukaryotic DNA polymerase processivity factor PCNA. Cell 79: 1233- 1243. Lai, C.J. and D. Nathans. 1975. Non-specific termination of simian virus 40 DNA rep1ication.J. Mol. Biol. 97: 113-118. Learn, B., A.W. Karzai, and R. McMacken. 1993. Transcription stimulates the establishment of bidirectional h DNA replication in vitro. Cold Spring Harbor Symp. Quant. Biol. 58: 389-402. Lee, E.H., A. Kornberg, M. Hidaka, T. Kobayashi, and T. Horiuchi. 1989. Escherichia coli replication termination protein impedes the action of helicases. Proc. Natl. Acad. Sci. 86: 9104-9108. Li, J.J. and R.J. Deshaies. 1993. Exercising self-restraint: Discouraging illicit acts of S and M in eukaryotes. Cell 74: 223-226.

.

458

B. Stillman

Liang, C., M. Weinreich, and B. Stillman. 1995. ORC and Cdc6p interact and determine the frequency of initiation of DNA replication in the genome. Cell 81: 667-676. Lin, S. and D. Kowalski. 1994. DNA helical instability facilitates initiation at the SV40 replication origin. J. Mol. Biol. 235: 496-507. Little, R.D. and C.L. Schildkraut. 1995. Initiation of latent DNA replication in the Epstein-Barr virus genome can occur at sites other than the genetically defined origin. Mol. Cell. Biol. 15: 2893-2903. Marahrens, Y. and B. Stillman. 1992. A yeast chromosomal origin of DNA replication defined by multiple functional elements. Science 255: 817-823. Marczynski, G.T., K. Lentine, and L. Shapiro. 1995. A developmentally regulated chromosomal origin of replication uses essential transcription elements. Genes Dev, 9: 1543-1557. Marians, K.J. 1992. Prokaryotic DNA replication. Annu. Rev. Biochem. 61: 673-719. McVey, D., S. Ray, Y. Gluzman, L. Berger, A.G. Wildeman, D.R. Marshak, and P. Tegtmeyer. 1993. cdc2 phosphorylation of threonine 124 activates the originunwinding functions of simian virus 40 T antigen. J. Virol. 67: 5206-5215. Melendy, T. and B. Stillman. 1993. An interaction between replication protein A and SV40 T antigen appears essential for primosome assembly during SV40 DNA replication. J. Biol. Chem. 268: 3389-3395. Mensa-Wilmot, K., K. Carroll, and R. McMacken. 1989. Transcriptional activation of bacteriophage h DNA replication in vitro: Regulatory role of histone-like protein HU of Escherichia coli. EMBO J. 8: 2393-2402. Messer, W., M. Meijer, H.E.N. Bergmans, F.G. Hanse, K. von Meyenburg, E. Beck, and H. Schaller. 1979. Origin of replication, oriC, of the Escherichia coli K12 chromosome: Nucleotide sequence. Cold Spring Harbor Symp. Quunt. Biol. 43: 139- 145. Morrical, S.W., K. Hempstead, and M.D. Morrical. 1994. The gene 59 protein of bacteriophage T4 modulates the intrinsic and single-stranded DNA-stimulated ATPase activities of gene 41 protein, the T4 replicative DNA helicase. J. Biol. Chem. 269: 33069-33081. Murray, A. 1994. Cell cycle checkpoints. Curr. Opin. in Cell Biol. 6: 872-876. Natale, D.A., R.M. Umek, and D. Kowalski. 1993. Ease of DNA unwinding is a conserved property of yeast replication origins. Nucleic Acids Res. 21: 555-560. Nethanel, T. and G. Kaufmann. 1990. Two DNA polymerases may be required for synthesis of the lagging DNA strand of simian virus 40. J. Virol. 64: 5912-5918. Nichols, A.F. and A. Sancar. 1992. Purification of PCNA as a nucleotide excision repair protein. Nucleic Acids Res. 20: 2441-2446. Nurse, P. 1994. Ordering S phase and M phase in the cell cycle. Cell 79: 547-550. O’Donnell, M., R. Onrust, F.B. Dean, M. Chen, and J. Hurwitz. 1993. Homology in accessory proteins of replicative polymerase-E. coli to humans. Nucleic Acids Res. 21: 1-3. Parsons, C.A., A. Stasiak, R.J. Bennett, and S.C. West. 1995. Structure of a multisubunit complex that promotes DNA branch migration. Nature 374: 375-378. Parsons, R., M.E. Anderson, and P. Tegtmeyer. 1990. Three domains in the simian virus 40 core origin orchestrate the binding, melting, and DNA helicase activities of T antigen. J. Virol. 64: 509-518. Rao, H. and B. Stillman. 1995. The origin recognition complex interacts with a bipartitie DNA binding site within yeast replicators. Proc. Nutl. Acad. Sci. 92: 2224-2228.

Prokaryotic and Eukaryotic DNA Replication

459

Roca, J. 1995. The mechanism of DNA topoisomerases. Trends Biochem. Sci. 20: 156-160. Rowley, A., J.H. Cocker, J. Hanvood, and J.F.X. Diffley. 1995. Initiation complex assembly at budding yeast replication origins begins with the recognition of a bipartitie sequence by limiting amounts of the initiator, ORC. EMBU J. 14: 2631-2641. Seki, M., T. Kohda, T. Yano, S. Tada, J. Yanagisawa, T. Eki, M. Ui, and T. Enomoto. 1995. Characterization of DNA synthesis and DNA-dependent ATPase activity at a restrictive temperature in temperature-sensitive tsFT848 cells with thermolabile DNA helicase B. Mol. Cell. Biol. 15: 165-172. Sekimizu, K. and A. Kornberg. 1988. Cardiolipin activation of dna A protein, the initiation protein of replication in Escherichia coli. J. Biol. Chem. 263: 7131-7135. Sekimizu, K., D. Bramhill, and A. Kornberg. 1988a. Sequential early stages in the in vitro initiation of replication at the origin of the Escherichia coli chromosome. J. Biol. Chem. 263: 7124-7130. Sekimizu, K., B.Y. Yung, and A. Kornberg. 1988b. The dna A protein of Escherichia coli: Abundance, improved purification and membrane binding. J. Biol. Chem. 263: 7136-7140. Shivji, M.K.K., M.K. Kenny, and R.D. Wood. 1992. Proliferating cell nuclear antigen is required for DNA excision repair. Cell 69: 367-374. Stahl, H., P. Droge, and R. Knippers. 1986. DNA helicase activity of SV40 large tumor antigen. EMBU J. 5 : 1939-1944. Stillman, B. 1989. Initiation of eukaryotic DNA replication in vitro. Annu. Rev. Cell. Biol. 5: 197-245. . 1994a. Initiation of chromosomal DNA replication in eukaryotes. Lessons from lambda. J. Biol. Chem. 269: 7047-7050. -. 1994b. Smart machines at the DNA replication fork. Cell 78: 725-728. Stukenberg, P.T., J. Turner, and M. O’Donell. 1994. An explanation for lagging strand replication: Polymerase hopping among DNA sliding clamps. Cell 78: 877-887. Taylor, J.H. 1977. Increase in DNA replication sites in cells held at the beginning of S phase. Chromosoma 62: 291-300. Tsurimoto, T. and B. Stillman. 1990. Functions of replication factor C and proliferating cell nuclear antigen: Functional similarity of DNA polymerase accessory proteins from human cells and bacteriophage T4. Proc. Natl. Acad. Sci. 87: 1023-1027. Tsurimoto, T., T. Melendy, and B. Stillman. 1990. Sequential initiation of lagging and leading strand synthesis by two DNA polymerase complexes at the SV40 DNA replication origin. Nature 346: 534-539. Ullsperger, C.J., A.V. Vologodskii, and N.R. Cozzarelli. 1995. Unlinking of DNA by topoisomerases during DNA replication. Nucleic Acids Mol. Biol. 9: 115-142. von Meyenburg, K., F.G. Hansen, E. Riise, H.E.N. Bergmans, M. Meijer, and W. Messer. 1979. Origin of replication, oriC, of the Escherichia coli K12 chromosome: Genetic mapping and minichromosome replication. Cold Spring Harbor Symp. Quant. Biol. 43: 121-1 28. Waga, S. and B. Stillman. 1994. Anatomy of a DNA replication fork revealed by reconstitution of SV40 DNA replication in vitro. Nature 369: 207-212. Wellinger, R.J. and V.A. Zakian. 1989. Lack of positional requirements for autonomously replicating sequence elements on artificial yeast chromosomes. Proc. Natl. Acad. Sci. 86: 973-977. Woese, C.R. and N.R. Pace. 1993. Probing RNA structure, function, and history by com-

460

9. Stillman

parative analysis. In The RNA world (ed. R.F. Gesteland and J.F. Atkins), pp. 91-117. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York. Ziegelin, G., E. Scherzinger, R. Lurz, and E. Lanka. 1993. Phage P4 a protein is multifunctional with origin recognition, helicase and primase activities. EMBO J. 12: 3703.

Suggest Documents